Next Article in Journal
Shedding the Light on Litopenaeus vannamei Differential Muscle and Hepatopancreas Immune Responses in White Spot Syndrome Virus (WSSV) Exposure
Next Article in Special Issue
Multiple FGF4 Retrocopies Recently Derived within Canids
Previous Article in Journal
Convergence of Prognostic Gene Signatures Suggests Underlying Mechanisms of Human Prostate Cancer Progression
Previous Article in Special Issue
Evolution of the T-Cell Receptor (TR) Loci in the Adaptive Immune Response: The Tale of the TRG Locus in Mammals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Characterization and Expression Analysis of NHX Gene Family under Salinity Stress in Gossypium barbadense and Its Comparison with Gossypium hirsutum

1
Biotechnology Research Institute, Chinese Academy of Agricultural Sciences, Beijing 100081, China
2
Genomics Lab, Department of Plant Breeding and Genetics, Faculty of Agricultural Science & Technology, Bahauddin Zakariya University, Multan 60000, Pakistan
3
Institute of Plant Breeding and Biotechnology, MNS-University of Agriculture, Multan 60000, Pakistan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2020, 11(7), 803; https://doi.org/10.3390/genes11070803
Submission received: 13 June 2020 / Revised: 6 July 2020 / Accepted: 14 July 2020 / Published: 16 July 2020
(This article belongs to the Special Issue A Tale of Genes and Genomes)

Abstract

:
Cotton is an important economic crop affected by different abiotic stresses at different developmental stages. Salinity limits the growth and productivity of crops worldwide. Na+/H+ antiporters play a key role during the plant development and in its tolerance to salt stress. The aim of the present study was a genome-wide characterization and expression pattern analysis under the salinity stress of the sodium-proton antiporter (NHX) of Gossypium barbadense in comparison with Gossypium hirsutum. In G. barbadense, 25 NHX genes were identified on the basis of the Na+_H+ exchanger domain. All except one of the G. barbadense NHX transporters have an Amiloride motif that is a known inhibitor of Na+ ions in plants. A phylogenetic analysis inferred three classes of GbNHX genes—viz., Vac (GbNHX1, 2 and 4), Endo (GbNHX6), and PM (GbNHX7). A high number of the stress-related cis-acting elements observed in promoters show their role in tolerance against abiotic stresses. The Ka/Ks values show that the majority of GbNHX genes are subjected to strong purifying selection under the course of evolution. To study the functional divergence of G. barbadense NHX transporters, the real-time gene expression was analyzed under salt stress in the root, stem, and leaf tissues. In G. barbadense, the expression was higher in the stem, while in G. hirsutum the leaf and root showed a high expression. Moreover, our results revealed that NHX2 homologues in both species have a high expression under salinity stress at higher time intervals, followed by NHX7. The protein-protein prediction study revealed that GbNHX7 is involved in the CBL-CIPK protein interaction pathway. Our study also provided valuable information explaining the molecular mechanism of Na+ transport for the further functional study of Gossypium NHX genes.

1. Introduction

Soil salinity is one of major abiotic stresses that limits crop production worldwide [1], with an estimated 45 million hectares of irrigated land reported to be under salinity stress. The world’s food production is mainly dependent on irrigated land, as it produces twice as much as the rain-fed area, therefore high salinity is a threat to sustainable crop production for the ever-increasing population [2,3]. By the year 2050, about 50% of all cultivable land is predicted to be affected by high salinization [4,5].
Most plants, being glycophytes, are affected by high levels of salt in the soil [6]. Plants have developed different mechanisms such as ionic stress pathways, oxidative stress pathways, and detoxification signalling to cope with the high soil salinity and toxicity of Na+ and Cl ions [7]. Many cellular processes conferring stress tolerance and regulating plant growth and development are dependent upon pH and ion homeostasis [8]. Ion-specific salinity is caused by the accumulation of toxic concentrations of sodium (Na+) and/or chloride (Cl) ions, especially in the older leaves [9]. In most plant species, the Na+ reaches the toxic concentration earlier than other salts [10]. Two non-selective cation channels (NSCC) are the major source of entry of Na+ into the cell; voltage-dependent and voltage-independent cation channels. The voltage-independent cation channels are thought to be a significant way of entering for Na+ ions. [11,12]. Sodium-hydrogen antiporters (NHX) are important antiporter genes which can help plants to exclude Na+ and Cl- ions through membranes or deposits in the vacuole to maintain the cell osmotic level [13]. Vacuole-bounded NHX antiporters regulate pH by countering acidity due to H+ pumps and functions such as H+ leaks to maintain the pH [14]. Besides the compartmentalization of Na+, NHXs could play a role in increasing the salinity tolerance by adjusting the K+ homeostasis [15,16,17].
Sodium-hydrogen antiporters (NHX) belongs to the cation proton antiporter1 (CPA1) family, which seems to have evolved from the sodium-proton antiporter (NhaP) genes in prokaryotes [18,19,20]. Human HsNHE was the first eukaryotic sodium hydrogen exchanger gene to be identified [21]. Meanwhile, in plants NHX1 was the first sodium hydrogen exchanger identified in Arabidopsis [22]. Besides contributing to salt tolerance [23], NHXs have diverse roles in biochemical and physiological processes, which include maintaining the pH in flowers [24], cellular expansion [25], K+ homeostasis [26], protein targeting, and vesicular trafficking [19,27,28]. Arabidopsis have eight members of the NHX genes that are further categorised into three groups based on their location. AtNHX1-4 is located in the vacuolar membrane, AtNHX7 and AtNHX8 are located in the plasma membrane, while AtNHX5 and AtNHX6 are located in the endosomal compartments [29,30]. The plasma membrane-bounded activity of the Na+/H+ antiporter activity has been studied in barley [31], tomato [32], and wheat [33], while a tonoplast-associated Na+/H+-antiporter activity has been reported for sugar beet [34], barley [35], sunflower [36], and Arabidopsis [23]. In Arabidopsis, the Na+ ion efflux is processed by the plasma membrane located Na+/H+ antiporter AtSOS1 under high salinity [37], while the vacuolar Na+/H+ antiporter catalyzes the sequestration of Na+ in vacuoles. Different studies have shown that the over-expression of NHX1 enhanced the plant tolerance towards salinity in different crops [20,23,38,39,40]; wheat NHX2 (TaNHX) transformed into alfalfa enhanced the salinity tolerance due to the homeostasis of potassium [41], whereas the nhx5 nhx6 double-knockout mutant in Arabidopsis aborted the transport through the tonoplast, increasing the sensitivity to salt stress [29]. These studies provide convincing proof of the involvement of the NHX genes in salinity tolerance, and this can be further explored in economically important crops.
Cotton is a worldwide leading textile fiber crop that has a significant impact on the economy of many agricultural-based countries [42]. G. barbadense and G. hirsutum, the two allotetraploids, are the most widely cultivated cotton species. With drastic environmental changes leading to a decline in the cultivated land area, like many other economic crops, cotton planting fields are moving to salinity and drought-affected areas. Overall, cotton crop production is always hindered by abiotic stresses, such as cold, heat, drought, and salinity [43,44]. Despite the fact that there are some natural varieties that are tolerant to drought and salinity, most high-quality cotton cultivars are sensitive to drought and salinity; in those cultivars, high soil salt concentrations affect the germination and emergence of seedlings [45,46], root growth [47,48], flowering, boll development, and fiber quality [49,50,51], causing an up to 50% loss in yield [52]. Finding the mechanism of abiotic stress tolerance will be of great significance for cotton production and genetic improvement.
In this study, we performed a genome-wide analysis of NHX genes in G. barbadense in comparison with G. hirsutum, including the phylogenetic relationships, a motif analysis, promoter analysis, the transcript expression under salt stress in different tissues, the chromosomal location, and the gene structures. The sequencing of many cotton species provides a wide range of genome data resources for gene family research [53,54,55,56,57,58]. Through a systematic analysis of all the members of the NHX gene, we can compare the gene regulation, expression pattern, and eventually their biological functions in cotton.

2. Materials and Methods

2.1. Characterization of Sodium Proton Antiporters

The NHX transporters are characterized by an Na+_H+_Exchanger domain (PF00999) (http://pfam.xfam.org/) [59]. The amino acid sequences of the NHX genes of G. hirsutum (JGI Version 2.0), G. barbadense (HAU, Version 1.0), G. arboreum (CRI, Version 1.0), and G. raimondi (JGI, Version 2.0) were downloaded from CottonFGD (https://cottonfgd.org/) [60] and were scanned against the Na+_H+_Exchanger domain using the HMMER 3.1b2 online software (https://www.ebi.ac.uk/Tools/hmmer/) [61]. The transmembrane domain prediction was made using the TMHMM server v.2.0 (http://www.cbs.dtu.dk/services/TMHMM/) [62], and, for the subcellular localization, CELLO v.2.5 (http://cello.life.nctu.edu.tw/) [63,64] was used. Netphos 3.1 [65] was used for the phosphorylation sites, and the prediction of the conserved motifs was carried out by MEME [66] with parameters such as the 2–20 motif sites, 10 no. of motifs, and 6–50-wide motif width; for all other tools, the default settings were used. The molecular weight (MW) and isoelectric point (pI) of the amino acid sequences were predicted using the online program ProtParam (http://web.expasy.org/protparam/).

2.2. Phylogeny and Divergence Analysis

A maximum likelihood phylogenetic tree was constructed with the amino acid sequences of Gossypium hirsutum (Gh), Gossypium barbadence (Gb), Gossypium arboreum (Ga), Gossypium raimondii (Gr), Arabidopsis thaliana (At), Vitis vinifera (Vv), Poplus trichocarpa (Ptr), Sorghum bicolor (Sb), Medicago truncatula (Mt), Eutrema halophilum (Eh), and Physcomitrella patens (Pp). The NHX protein sequences of four cotton species were downloaded from Cotton FGD and were already reported for S. bicolor and P. patens [67]. T. halophilum, also known as E. halophilum [68]; V. vinifera; P. trichocarpa [69]; A. thaliana [37]; and M. truncatula [70] were downloaded from the online Phytozome v11 (https://phytozome.jgi.doe.gov/pz/portal.html) (Table S1). All the retrieved amino acid sequences were confirmed with the hidden Markov model (HMM) using the PF00999 Na+_H+ exchanger domain. The sequences were then aligned using muscle and subjected to a phylogenetic analysis using MEGA 10.0 [71]; the bootstrap value was kept at 1000. The resulting tree was visualized using iTOL v5 (https://itol.embl.de/) [72]. Tbtools [73] were used to estimate the gene duplication events. To further calculate the synonymous (ds) and non-synonymous (dN) substitution rates, the PAL2NAL program [74] was used.

2.3. Promoter and Gene Structure Analysis

The upstream 2 Kb sequences for all the NHX genes of G. barbadense and G. hirsutum were analyzed in silico to find out the potential Cis-acting elements. All the promoters were submitted to PLANTCARE [75], and the resulting Cis-acting elements were categorized based on their functional class. The Gene Structure Display Server tool was used for the analysis of the gene structure [76].

2.4. Protein-Protein Interaction and Physical Mapping

The STRING database (https://string-db.org) was used to predict the protein-protein interactions. The genomic coordinates of the transporters were extracted from the Cotton FGD Database (https://cottonfgd.org/) [60] using the HAU assembly for G. barbadense and the JGI assembly for G. hirsutum and then used to map the genes onto different chromosomes physically.

2.5. Expression Analysis under Salinity

To investigate the expression level of NHX transporters under salinity stress, the G. barbadense cultivar Hai7124 and the G. hirsutum cultivar TM-1 were sown in pots in greenhouse conditions with temperatures ranging from 25 to 30 °C and with 12 h light and 12 h dark. At the emergence of true leaves, the seedlings were treated with a 400 mM salinity level and tap water served as a control. Samples were taken from the leaves, stems, and roots 0 h, 3 h, 6h, and 12 h after the treatment, then they were snap-frozen in liquid nitrogen and subsequently stored at −80 °C until the RNA was extracted.

2.6. RNA Extraction and Quantitative Real-Time PCR Analysis

The total RNA was isolated from all the samples using the EASYspin RNA plant-kit (Cat#DR103-03) following the instruction manual. DNaseI (RNase-free) was used to eliminate the genomic DNA contamination in the RNA samples. The concentration and purity was checked by Thermo fisher Scientific Nano-Drop One and run on 1% agarose gel. The total RNA (5 g) was taken as a template for a first strand cDNA synthesis using the iScriptTm Reverse Transcription Supermix for RT-qPCR (BIO-RAD, Hercules, CA, USA).
BIO-RAD’s CFX Connect Real-Time PCR Detection System was used to study the relative expression level of the G.barbadense and G. hirsutum NHX genes using the iTAQ UNIVERSAL SYBR GREEN MIX (BIO-RAD) with gene-specific primers. Each gene expression was normalized with the Actin genes [77]. The thermal cycler conditions were 95 °C for 3 min, followed by 40 cycles of 95 °C for 10 s, 60 °C for 1 s, and 72 °C for 30 s, and the melting curve stage was at 95 °C for 10 s, 65 °C for 1 min, and 97 °C for 1–5 s.

3. Results

3.1. Characterization of NHX Genes in Cotton Species

To retrieve the members of the NHX gene family, we searched four cotton species’ genome data based on the Na+_H+_Exchanger domain (PF00999). A total of 25 NHX genes in G. barbadense, 23 in G. hirsutum, 13 in G. arboreum, and 13 in G. raimondii were identified. We further determined the biophysical properties of the G. barbadense NHX genes including the locus ID, CDS length (bp), protein length (aa), Na+/H+ exchanger domain, predicted protein molecular weight (MW), predicted cellular localization, isoelectric points (pI), and trans-membrane domains. The NHX proteins were predicted to be localized on the plasma membrane, endoplasmic reticulum, and vacuole with number of amino acids ranging from 164 to 1152. The molecular weight ranges from 18.66 kDa (Gb-NHX1) to 128.14 kDa (Gb-NHX7-1D) (Table 1, Figure S1). Previously, it has been reported that the distribution pattern of intron/exon in a gene play a vital role in its biological function. The number of exons for both the G. barbadense and G. hirsutum transporters varies from 14 to 23, with the exception of NHX1 (Figure 1). Moreover, the gene structure analysis of the tetraploid cotton species (G. hirsutum and G. barbadense) along with the phylogeny results showed that the genes with a similar intron/exon pattern clustered near to each other in same groups (Figure S2). An in silico analysis revealed that the NHX transporters are mostly phosphorylated with protein kinase C, cyclin-dependent protein kinase (CDC2), and protein kinase A (PKA), respectively, and very less with the ataxia telangiectasia mutated (ATM). The most common site for phosphorylation was serine, in comparison with theorine and tyrosine (Table S2).

3.2. Phylogeny and Sequence Logos of GbNHX Genes with Different Species

In order to find the evolutionary relationship among the NHX genes, the protein sequences from 11 different plant species, including 4 gossypium species, G. hirsutum, G. barbadence, G. arboreum, and G. raimondii; 5 dicotyledonous angiosperms, A. thaliana, V. vinifera, P. trichocarpa, M. truncatula, and E. halophilum; one monocotyledonous angiosperm, S. bicolor; and one bryophyte, P. patens, were retrieved. A maximum likelihood tree was constructed among 123 NHX genes of the above-mentioned plant species. The phylogenetic tree depicted a direct relation with the subcellular localization, as all the NHX transporter proteins from different species clustered in three clades based upon their predicted location—viz., VAC (vacuolar membrane-bounded), ENDO (endomembrane-bounded), and PM (plasma membrane-bounded). Moreover, the VAC class has 85 genes, as most types of NHX genes (NHX1, 2, 3, and 4) from different species are present on the vacuolar membrane, while ENDO has 20 and the PM class has 18 genes. Among the gossypium species, the VAC class has NHX1, NHX2, and NHX4; the ENDO class has NHX6; and the PM class has NHX7 (Figure 2). To investigate the amino acid changes in the NHX domain across four cotton species, we generated the sequence logos of conserved amino acids. We found that many sequence logos were highly conserved across the N and C termini among different species. Within a species, the NHX domain of G. raimondii has the most conserved sequences (Figure S3).

3.3. Comparison of Motifs and Physical Genome Mapping of NHX Genes in G. barbadense and G. hirsutum

A motif prediction carried out by MEME with 0–10 motif sites showed that all of the G. barbadense NHX transporters except one (Gb-NHX2-2A) have an amiloride binding motif, while in the case of G. hirsutum, all transporters have this motif (Figure 3a,b). To further investigate the presence of this motif in the NHX genes of other species, we aligned 99 amino acid sequences from the gossypium species, V. vinifera, M. trunculata, A. thaliana, and P. trichocarpa. Our results showed that almost all (97) the NHX transporters have an amiloride binding site, except Gb-NHX2-2A and GaNHX6-1 of G. barbadense and G. arboreum, respectively (Figure S4). The physical mapping of the NHX transporters on the corresponding chromosomal loci in four Gossypium species showed that the NHX genes are scattered on both the A and D genomes. In G. barbadense, 12 genes were mapped on the At sub-genome, while 13 were mapped on the Dt sub-genome. In case of G. hirsutum, the At sub genome has 11 and the Dt sub-genome has 12 NHX genes. In both the allotertaploid species, A01, A09, A11, D01, D02, D09, and D11 have two, while A02, A03, A06, A12, A13, D06, D07, D12, and D13 have one NHX transporter each. Chromosomal mapping also showed some differences among both species, with only G. barbadense having one transporter on A08 and D08. Moreover, in Gb and Gh two NHX transporters were present on the chromosomes D01, DO9, and D11 each, while G. raimondii, the progenitor of the D genome, has no member on these chromosomes (Table 2, Figure S5).

3.4. Synteny Analysis and Ka/Ks Ratio of NHX in Cotton Species

To investigate the relationship among allotetraploid G. barbadense and its diploid ancestors G. arboreum and G. raimondii, a neighbor-end joining tree was constructed (Figure S6). The clusters formed in the tree with the same type of NHX genes from all three species provide evidence that G. barbadense is the result of hybridization between the two diploid cotton species, G. arboreum and G. raimondii.
Being an allotetraploid, upland cotton is a model crop species to study natural polyploidy [78]. To study the relationship between the GbNHX and GhNHX genes, orthologous/paralogous genes pairs were identified for the At and Dt sub-genomes. In accordance with previous findings, our study also demonstrated that the At as well as the Dt sub-genomes have orthologs in the A (G. arboreum) or D (G. raimondii) genomes (Figure 4a,b). The synteny analysis showed a total of 30 gene duplication events in G.barbadense, while there were 31 in G. hirsutum on the basis of a whole-genome analysis (Table 3). Most of the GbNHX genes showed whole-genome or segmental duplication. Furthermore, to estimate the selection pressure on the Gossypium NHX transporters during the evolutionary time, we calculated the Ka and Ks values and Ka/Ks ratio in both tetraploid species. The Ka/Ks ratio for most of the genes was less than 1, while for only three (Gb-NHX2-2A, Gb-NHX2-7D, and Gh_NHX6-3D) was it more than 1 (Table S4). This indicates that the cotton NHX genes have been subjected to strong purifying selection. Interestingly, an expression analysis also revealed that G. barbadense Gb-NHX2-7D and G. hirsutum Gh-NHX6-3D have a higher expression in different tissues under salinity stress.

3.5. Promoter Analysis of G. barbadense and G. hirsutum NHX Genes

Cis-acting elements in the promoter region play a key role in defining the plant response towards stress and light and in growth regulation. To investigate the transcriptional potential of the Na+/H+ transporter genes, we analyzed and predicted the Cis-elements in 2000 bp promoter regions upstream of the start codon. Besides the abundant amount of core promoter/enhancer elements—i.e., CAAT-Box (CAAT, CAAAT, and TGCCAAC) and TATA-box (ATTATA, TAAAGATT, TATTTAAA, TATA, ccTATAAAaa, TATACA), with a total number of 806 and 1178, respectively—we found different elements related to stress, light, and hormone response. Interestingly, the NHX genes contained a larger number of Cis-elements related to stress response than to light and hormone response, indicating their role in stress regulation. The water and drought response elements MYB (CAACCA/TAAC/TAACTG) and MYC (CAATTG/TCTCTTA/TCTCTTA) were the most abundant among all the elements present, with a total number of 89 (12%) and 72 (10%), respectively (Table S5). In G. barbadense, 21 GbNHXs have AREs (anaerobic-responsive elements); 17 have STREs (stress-responsive elements); 10 contained the WUN-motif (wound-response element); and 9 GbNHXs had a W-box, which is involved in pathogen response [79]. Meanwhile, the G. hirsutum NHXs have comparatively less putative Cis-elements, with 17 GhNHXs having AREs and 15 having STREs, while the WUN-motif and W-box were found in 9 and 7 GhNHXs, respectively. The promoter region of Gb-NHX7-1A (Gbar_A03G012870) and Gb-NHX7-1D (Gbar_D02G014810) has a maximum number of stress-responsive Cis-elements (Figure 5).

3.6. Expression Pattern of G. barbadense NHX Genes and Its Comparison with G. hirsutum under Salt Stress

The expression pattern of NHX genes under salinity stress was checked to investigate their potential role in G. barbadense and was compared to that of G. hirsutum. Previously, G. barbadense was found to be more tolerant to salinity than G. hirsutum [80,81]; studies showed that it has more lateral roots under a stress environment [82]. We used qRT-PCR for the expression analysis of all the NHX transporters in G.barbadense and G. hirsutum in the root, stem, and leaf tissue at 0, 3, 6, and 12 h time intervals. Our results revealed that in case of G. barbadesne, most genes show a higher expression level in the stem tissue, while in G. hirsutum, more genes are expressed in the roots and leaves, with a less significant expression in the stem under stress as compared with the control (Figure 6 and Figure 7). Ten GbNHX genes—Gb-NHX2-4A, Gb-NHX2-7A Gb-NHX2-8A, Gb-NHX7-1A, Gb-NHX2-2D, Gb-NHX2-3D, Gb-NHX2-7D, Gb-NHX2-8D Gb-NHX6-1D and Gb-NHX7-1D—with a higher expression were further analyzed (Figure 7). The genes showed differential expressions in different tissues. Almost all the genes showed a maximum expression at 12h in different tissues. Our results also showed that Gb-NHX2-7A, Gb-NHX2-3D, and Gb-NHX2-7D have a higher number of stress-related Cis-elements in their promoter region that could be related to high expression under stress. Additionally, the Ka/ks ratio revealed that Gb-NHX2-7D underwent positive selection. Moreover, we observed that the NHX2 homologues in both species have a high expression under salinity stress at higher time intervals, followed by NHX7. In G. barbadense, the plasma membrane-bounded NHX7 has a high expression level in all tissues under stress.

3.7. Protein-Protein Interaction Prediction and GO of GbNHX Genes

On the string database, only the Gossypium raimondii (Gr) protein-protein interaction network was available until now. Thus, in this study we used the homolog gene between GrNHX and GbNHX to search in the database. The GrNHX homolog gene and interacted protein were used to construct a network to predict the GbNHX protein-protein interaction network. We observed that the Gossypium NHX proteins interact with other proteins, such as HKT1, conferring salinity tolerance and RCD 1 (Radical-Induced Cell Death protein 1), which supports chloroplasts against high ROS (Reactive oxygen species). The NHX protein also interacted with calcineruin B-like proteins (CBL10) and some CBL-interacting protein kinases (CIPKs), such as CIPK8 and CIPK 24. Meanwhile, NHX7/SOS1 and SOS2, interacting with almost all proteins, were found to be the centers of interaction (Figure 8). When single proteins were subjected to analysis individually, they showed a similar kind of interaction with related proteins involved in stress tolerance (Figure S7). Moreover, the gene ontology (GO) of the GbNHX gene showed that they are enriched in 11 GO terms related to potassium ion homeostasis, the response to salt; the regulation of pH; sodium: proton antiporter activity; solute: proton antiporter activity; cation transport; transmembrane transport; the integral component of membrane; sodium ion transport; the vacuolar membrane; and the plasma membrane (Figure 9, Figure S8, Table S6).

4. Discussion

Salinity causes ion toxicity and physiological drought, thus limiting the growth and productivity of plants [2]. Recently, the availability of high-quality de novo genome assemblies for G. arboreum [56] and allotetraploids cottons [83] generate new opportunities for precise genome-wide studies in cotton. The NHXs genes present in plant cells maintain the ionic homeostasis by playing their role in the extrusion of Na+ ions out of the cell and the compartmentalization of Na+ ions into the vacuole [84]. In the current study, a total of 25 with different types—i.e., NHX1, NHX2, NHX4, NHX6, and NHX7—of sodium transporters have been identified in G. barbadense, based on the Na+_H+_Exchanger domain (Table 1).
A bioinformatics analysis showed that the NHX members in G. barbadense can be divided into three categories depending upon their subcellular location, with NHX7 localized in the plasma membrane, NHX6 in the endomembrane, and the others in the tonoplast. In Arabidopsis, both NHX7 and NHX8 are localized in the plasma membrane [85], while NHX5 and NHX6 are present in the endomembrane [29]. However, no NHX5 and NHX8 were observed in the Gossypium species in this study (Figure 2). Subcellular localization could be a key factor in defining the function of NXH transporters. NHX members located on both the plasma membrane and tonoplast play their role in the exclusion and compartmentalization of excess Na+ and maintain ionic homeostasis. Moreover, some NHX members that are endomembrane-bounded were found to be vital for cellular cargo trafficking, growth development, and the regulation of protein processing [13,29]. The phylogenetic analysis indicated that GbNHX has paralogous or orthologous groups with other Gossypium species members. The NHX genes in P. trichocarpa [69], S. bicolor [67], and B. vulgaris [86] showed three phylogenetic clusters based on their location in the cell; we found the same results for cotton NHX transporters. An amiloride binding site (L/F)FF(I/L)(Y/F)LLPPI, a typical feature of NHX transporters in plants [87,88], is present in the N-terminal of these proteins; the presence of amiloride even in a micro amount in the Na+/H+ exchangers inhibits the transport of Na+ transport [89]. This site was found in most of G. barbadense transporters, such as Arabidopsis and poplar [69] (Figure S2).
During the cotton evolution period, the occurrence of a gene duplication event led to the creation of new genes [90]. The origin of multi-gene families has been attributed to a region-specific gene duplication that occurred in upland cotton [53]. The presence of two or more genes on the same chromosome reveal the possibility of a tandem duplication event, while the genes present on different chromosomes result in a segmental duplication event. The duplication of genes increase the functional divergence, which is an essential factor in adoptability under changing environmental conditions [91]. The Ka/Ks ratio is a measure used to examine the mechanisms of gene duplication evolution after divergence from their ancestors. The Ka/Ks ratio gives an insight into the selection pressure on amino acid substitutions, with a Ka/Ks ratio < 1 indicating a purifying selection, while a ratio > 1 suggests the possibility of positive selection. Wang et al. [92] showed in T. aestivum and TaBT1 that the positive selection of a gene during evolution increases its potential and has more transcription levels under stress conditions. Almost all except 3 out of 31 duplication events occurred in the G.barbadense and G. hirsutum NHX transporters showing a <1 substitution value, indicating that these genes underwent a positive Darwinism or purifying selection [93] (Table 3).
The promoter region of G. barbadense and G. hirsutum NHX transporters has light, stress, and hormone- and development-responsive Cis-acting elements, showing that these genes are not only regulated by abiotic stress but also by different hormones (Table S5). However, the number of stress-responsive Cis-elements exceeds the others, indicating their major role in abiotic stress response (Figure 5). Similar to Arabidopsis [67], abscisic acid-responsive elements (ABRE), auxin-responsive elements, fungal-responsive elements, circadian elements, low temperature-responsive elements (LTR), heat shock elements (HSE), and MYB Cis-elements were noticed in the Gossypium barbadesne NHX gene promoter. The β-glucoronidase gene driven by the AtMYB2 promoter in Arabidopsis was found to be inducible by osmotic stresses [94]. G-box elements that act as positive regulators of early leaf senescence in rice [95] were also detected in the promoter regions of Gossypium NHX transporters, implying that these genes also modulate the leaf senescence.
In plants, sodium-proton antiporters facilitate both Na+/H+ and K+/H+ exchanges, therefore contributing to both stress tolerance and K+ nutrition [25,26,96]. NHX genes have been reported to enhance salinity tolerance in different species, such as A. thaliana [37], B. vulgaris [97], S, lycopersicum [40,98], H. vulgare [99], Z. maize [100], T. aestivum [101], G. max [102], O. sativa [103,104], and S. bicolor [67]. Our study revealed that in G. barbadense and G. hirsutum, the NHX genes express differentially in different tissues at different time intervals under salinity stress. Ma et al. [105] also observed different expression levels of NHX genes in different tissues of V. vinifera L. The vac-class NHX2 homologues in cotton show a higher expression under salinity stress. When R. trigyna is exposed to salinity stress, an increase in the transcription level of the vac-class RtNHX1 gene in leaves was observed [106]. A similar kind of expression pattern was observed in sweet potato, IBNHX2 [107] and in T. aestivum, TaNHX3 [108] under the salt treatment.
The plasma membrane-bounded NHX7/SOS1 gene helps in the exclusion of Na+ ions from the cell to regulate ionic homeostasis [5,109]; it was validated in the present study that Gb-NHX7 showed a higher expression under the salinity stress. It is noticeable that its expression is higher in roots at all time periods than in other tissues. Similar results have been noticed in Salicornia brachiate [110], P. tenuiflora [111], and Z. xanthoxylum [112], where plasma membrane-bounded NHX7/SOS1 showed a higher expression in roots than in shoots and was further increased at a higher salt stress. These results proposed that GbNHX7 could be responsible for the long distance transport of Na+ ions, but the detailed mechanism is still to be explored.
The protein-protein interaction showed that GbNHX interacted with many other proteins. The tails on the C-terminal of SOS1 and NHX1 were revealed to be essential for protein-protein interaction by Quintero at el. [113]. In Arabidopsis, SOS1 interacts with RCD1 (radical-induced cell death protein 1) to increase the tolerance against oxidative stress caused by ROS [114]. Our hypothesis also indicated the presence of interaction between NHX7/SOS1 and RCD1 to improve the salt tolerance ability of the plants. Moreover, in the present study HKT1 was found to interact with almost all the GbNHX genes. Zhang et al. [111] observed that under considerably high salt concentrations when vacuoles have no more capacity to sequester Na+ ions, the HKT1;5 is strongly expressed to increase the salinity tolerance by unloading excess Na+ ions from the xylem. The interaction between the CBL proteins and CIPK is also known to be involved in enhancing the ability of the plant to withstand salt stress [115]. Kim et al. [116] observed that CIPK24/SOS2 make a complex with CBL3/SOS3 that phosphorylates the NHX7/SOS1 localized in the plasma membrane to pump Na+ ions out from the cell. The single protein-protein interaction is this study also infers a similar kind of results, showing the interaction of NHX7/SOS1 with CIPK24 and CIPK8, besides others.

5. Conclusions

A genome-wide study of G. Barbadense revealed the presence of four types (NHX2, NHX4, NHX6, and NHX7) of sodium transporters that can be categorized as plasma membrane (GbNHX7), endomembrane (GbNHX6), and vacuolar (GbNHX1, 2, and 4), based on their location. The amiloride-binding site (FFIYLLPPI) is found in all GbNHX genes. The high number of stress related Cis-acting elements observed in promoters show their role in tolerance against abiotic stresses. A chromosomal localization and collinearity analysis showed the purified selection and evolution of gossypium NHX genes. An in silico PPI network analysis showed that only GbNHX7 interacts with CBLs and CIPKs, suggesting this protein might be the primary NHX involved in the CBL-CIPK pathway during the salt stress response. The gene ontology (GO) showed that these genes are involved in the proton antiport, sodium ion transport across the membrane, and salinity response activities. A tissue-specific qRT-PCR-based expression analysis of NHX antiporters revealed that they are more expressed under stress conditions in comparison with control conditions. The expression pattern was also different in different tissues of G. barbadense and G. hirsutum. The higher expression of vac-class in leaves may also be responsible for the deposition of salts, especially in older leaves. These results showed that these genes could be involved in various developmental processes and stress responses by maintaining the turgor pressure, pH, and ionic homeostasis. Our findings would be useful in selecting candidate genes for functional validation in relation to high soil salinity stress tolerance for the improvement of crop plants.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4425/11/7/803/s1: Table S1: Phylogeny Sequences. Table S2: Phosphorylation sites and gene rename file. Table S3. qRT-PCR primer sequences. Table S4: Ka/Ks ratio of Gb and Gh NHX genes. Table S5: Cis-acting elements for GBNHX and GhNHX. Table S6: Gene Ontologies term. Figure S1: Transmembrane Domains for GBNHX proteins. Figure S2: Gene Structure for GbNHX and GhNHX. Figure S3: Sequence logos of four cotton species. Figure S4: Amiloride binding site in NHX genes. Figure S5: Chromosomal map four for Gossypium species. Figure S6: Gb, Ga and Gr NHX genes phylogenetic tree. Figure S7: Single protein-protein Interaction. Figure S8: GO terms graph.

Author Contributions

Conceptualization, U.A., C.L., Z.M. and R.Z.; Data curation, U.A. and Z.M.; Formal analysis, U.A., Y.S., A.A.M. and M.A. (Mubashir Abbas); Funding acquisition, R.Z.; Methodology, U.A., M.A.A. and Z.M.; Project administration, R.Z.; Resources, S.G.; Software, U.A., Y.S. and M.A. (Muhammad Askari); Supervision, R.Z.; Validation, M.A.A., Z.M. and R.Z.; Visualization, C.L.; Writing—original draft, U.A. and Y.S.; Writing—review and editing, W.M., Z.A., Z.M. and R.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded National Key R&D Program of China (Grant Number 2016YFE0117600).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Quesada, V.; Ponce, M.R.; Micol, J.L. Genetic analysis of salt-tolerant mutants in Arabidopsis thaliana. Genetics 2000, 154, 421–436. [Google Scholar] [PubMed]
  2. Munns, R.; Tester, M. Mechanisms of salinity tolerance. Annu. Rev. Plant Biol. 2008, 59, 651–681. [Google Scholar] [CrossRef] [Green Version]
  3. Shabala, S.; Cuin, T.A. Potassium transport and plant salt tolerance. Physiol. Plant. 2008, 133, 651–669. [Google Scholar] [CrossRef]
  4. Hasegawa, P.M. Sodium (Na+) homeostasis and salt tolerance of plants. Environ. Exp. Bot. 2013, 92, 19–31. [Google Scholar] [CrossRef]
  5. Zhu, J.K. Regulation of ion homeostasis under salt stress. Curr. Opin. Plant Biol. 2003, 6, 441–445. [Google Scholar] [CrossRef]
  6. Slama, I.; Abdelly, C.; Bouchereau, A.; Flowers, T.; Savoure, A. Diversity, distribution and roles of osmoprotective compounds accumulated in halophytes under abiotic stress. Ann. Bot. 2015, 115, 433–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Yang, Y.; Guo, Y. Unraveling salt stress signaling in plants. J. Integr. Plant Biol. 2018, 60, 796–804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. An, R.; Chen, Q.J.; Chai, M.F.; Lu, P.L.; Su, Z.; Qin, Z.X.; Chen, J.; Wang, X.C. AtNHX8, a member of the monovalent cation: Proton antiporter-1 family in Arabidopsis thaliana, encodes a putative Li+/H+ antiporter. Plant J. 2007, 49, 718–728. [Google Scholar] [CrossRef]
  9. Roy, S.J.; Negrão, S.; Tester, M. Salt resistant crop plants. Curr. Opin. Biotechnol. 2014, 26, 115–124. [Google Scholar] [CrossRef]
  10. Tester, M.; Davenport, R. Na+ tolerance and Na+ transport in higher plants. Ann. Bot. 2003, 91, 503–527. [Google Scholar] [CrossRef]
  11. Schachtman, D.; Liu, W. Molecular pieces to the puzzle of the interaction between potassium and sodium uptake in plants. Trends Plant Sci. 1999, 4, 281–287. [Google Scholar] [CrossRef]
  12. Amtmann, A.; Sanders, D. Mechanisms of Na+ uptake by plant cells. In Advances in Botanical Research; Elsevier: York, UK, 1998; Volume 29, pp. 75–112. [Google Scholar]
  13. Bassil, E.; Coku, A.; Blumwald, E. Cellular ion homeostasis: Emerging roles of intracellular NHX Na+/H+ antiporters in plant growth and development. J. Exp. Bot. 2012, 63, 5727–5740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Sze, H.; Chanroj, S. Plant endomembrane dynamics: Studies of K+/H+ antiporters provide insights on the effects of pH and ion homeostasis. Plant Physiol. 2018, 177, 875–895. [Google Scholar] [CrossRef] [Green Version]
  15. Wu, H.; Zhang, X.; Giraldo, J.P.; Shabala, S. It is not all about sodium: Revealing tissue specificity and signalling roles of potassium in plant responses to salt stress. Plant Soil 2018, 431, 1–17. [Google Scholar] [CrossRef]
  16. Ma, Y.C.; Augé, R.M.; Dong, C.; Cheng, Z.M. Increased salt tolerance with overexpression of cation/proton antiporter 1 genes: A meta-analysis. Plant Biotechnol. J. 2017, 15, 162–173. [Google Scholar] [CrossRef]
  17. Sharma, H.; Taneja, M.; Upadhyay, S.K. Identification, characterization and expression profiling of cation-proton antiporter superfamily in Triticum aestivum L. and functional analysis of TaNHX4-B. Genomics 2020, 112, 356–370. [Google Scholar] [CrossRef]
  18. Mäser, P.; Thomine, S.; Schroeder, J.I.; Ward, J.M.; Hirschi, K.; Sze, H.; Talke, I.N.; Amtmann, A.; Maathuis, F.J.; Sanders, D. Phylogenetic relationships within cation transporter families of Arabidopsis. Plant Physiol. 2001, 126, 1646–1667. [Google Scholar] [CrossRef] [Green Version]
  19. Brett, C.L.; Donowitz, M.; Rao, R. Evolutionary origins of eukaryotic sodium/proton exchangers. Am. J. Physiol. Cell Physiol. 2005, 288, C223–C239. [Google Scholar] [CrossRef] [Green Version]
  20. Rodríguez-Rosales, M.P.; Gálvez, F.J.; Huertas, R.; Aranda, M.N.; Baghour, M.; Cagnac, O.; Venema, K. Plant NHX cation/proton antiporters. Plant Signal. Behav. 2009, 4, 265–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Sardet, C.; Franchi, A.; Pouysségur, J. Molecular cloning, primary structure, and expression of the human growth factor-activatable Na+ H+ antiporter. Cell 1989, 56, 271–280. [Google Scholar] [CrossRef]
  22. Gaxiola, R.A.; Rao, R.; Sherman, A.; Grisafi, P.; Alper, S.L.; Fink, G.R. The Arabidopsis thaliana proton transporters, AtNhx1 and Avp1, can function in cation detoxification in yeast. Proc. Natl. Acad. Sci. USA 1999, 96, 1480–1485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Apse, M.P.; Aharon, G.S.; Snedden, W.A.; Blumwald, E. Salt tolerance conferred by overexpression of a vacuolar Na+/H+ antiport in Arabidopsis. Science 1999, 285, 1256–1258. [Google Scholar] [CrossRef] [PubMed]
  24. Yamaguchi, T.; Fukada-Tanaka, S.; Inagaki, Y.; Saito, N.; Yonekura-Sakakibara, K.; Tanaka, Y.; Kusumi, T.; Iida, S. Genes encoding the vacuolar Na+/H+ exchanger and flower coloration. Plant Cell Physiol. 2001, 42, 451–461. [Google Scholar] [CrossRef] [Green Version]
  25. Apse, M.P.; Sottosanto, J.B.; Blumwald, E. Vacuolar cation/H+ exchange, ion homeostasis, and leaf development are altered in a T-DNA insertional mutant of AtNHX1, the Arabidopsis vacuolar Na+/H+ antiporter. Plant J. 2003, 36, 229–239. [Google Scholar] [CrossRef] [PubMed]
  26. Leidi, E.O.; Barragán, V.; Rubio, L.; El-Hamdaoui, A.; Ruiz, M.T.; Cubero, B.; Fernández, J.A.; Bressan, R.A.; Hasegawa, P.M.; Quintero, F.J. The AtNHX1 exchanger mediates potassium compartmentation in vacuoles of transgenic tomato. Plant J. 2010, 61, 495–506. [Google Scholar] [CrossRef] [Green Version]
  27. Bowers, K.; Levi, B.P.; Patel, F.I.; Stevens, T.H. The sodium/proton exchanger Nhx1p is required for endosomal protein trafficking in the yeast Saccharomyces cerevisiae. Mol. Biol. Cell 2000, 11, 4277–4294. [Google Scholar] [CrossRef] [Green Version]
  28. Sottosanto, J.B.; Gelli, A.; Blumwald, E. DNA array analyses of Arabidopsis thaliana lacking a vacuolar Na+/H+ antiporter: Impact of AtNHX1 on gene expression. Plant J. 2004, 40, 752–771. [Google Scholar] [CrossRef] [Green Version]
  29. Bassil, E.; Ohto, M.-A.; Esumi, T.; Tajima, H.; Zhu, Z.; Cagnac, O.; Belmonte, M.; Peleg, Z.; Yamaguchi, T.; Blumwald, E. The Arabidopsis intracellular Na+/H+ antiporters NHX5 and NHX6 are endosome associated and necessary for plant growth and development. Plant Cell 2011, 23, 224–239. [Google Scholar] [CrossRef] [Green Version]
  30. Bassil, E.; Tajima, H.; Liang, Y.-C.; Ohto, M.-A.; Ushijima, K.; Nakano, R.; Esumi, T.; Coku, A.; Belmonte, M.; Blumwald, E. Correction: The Arabidopsis Na+/H+ Antiporters NHX1 and NHX2 Control Vacuolar pH and K+ Homeostasis to Regulate Growth, Flower Development, and Reproduction. Plant Cell 2011, 23, 4526. [Google Scholar] [CrossRef] [Green Version]
  31. Ratner, A.; Jacoby, B. Effect of K+, its counter anion, and pH on sodium efflux from barley root tips. J. Exp. Bot. 1976, 27, 843–852. [Google Scholar] [CrossRef]
  32. Mennen, H.; Jacoby, B.; Marschner, H. Is sodium proton antiport ubiquitous in plant cells? J. Plant Physiol. 1990, 137, 180–183. [Google Scholar] [CrossRef]
  33. Allen, G.J.; Sanders, D. Calcineurin, a type 2B protein phosphatase, modulates the Ca2+-permeable slow vacuolar ion channel of stomatal guard cells. Plant Cell 1995, 7, 1473–1483. [Google Scholar] [CrossRef] [PubMed]
  34. Blumwald, E.; Poole, R.J. Na+/H+ antiport in isolated tonoplast vesicles from storage tissue of Beta vulgaris. Plant Physiol. 1985, 78, 163–167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Garbarino, J.; DuPont, F.M. NaCl induces a Na+/H+ antiport in tonoplast vesicles from barley roots. Plant Physiol. 1988, 86, 231–236. [Google Scholar] [CrossRef] [Green Version]
  36. Ballesteros, E.; Blumwald, E.; Donaire, J.P.; Belver, A. Na+/H+ antiport activity in tonoplast vesicles isolated from sunflower roots induced by NaCl stress. Physiol. Plant. 1997, 99, 328–334. [Google Scholar] [CrossRef]
  37. Shi, H.; Ishitani, M.; Kim, C.; Zhu, J.-K. The Arabidopsis thaliana salt tolerance gene SOS1 encodes a putative Na+/H+ antiporter. Proc. Natl. Acad. Sci. USA 2000, 97, 6896–6901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Liu, H.; Wang, Q.; Yu, M.; Zhang, Y.; Wu, Y.; Zhang, H. Transgenic salt-tolerant sugar beet (Beta vulgaris L.) constitutively expressing an Arabidopsis thaliana vacuolar Na/H antiporter gene, AtNHX3, accumulates more soluble sugar but less salt in storage roots. Plant Cell Environ. 2008, 31, 1325–1334. [Google Scholar] [CrossRef]
  39. Ohta, M.; Hayashi, Y.; Nakashima, A.; Hamada, A.; Tanaka, A.; Nakamura, T.; Hayakawa, T. Introduction of a Na+/H+ antiporter gene from Atriplex gmelini confers salt tolerance to rice. FEBS Lett. 2002, 532, 279–282. [Google Scholar] [CrossRef] [Green Version]
  40. Zhang, H.X.; Blumwald, E. Transgenic salt-tolerant tomato plants accumulate salt in foliage but not in fruit. Nat. Biotechnol. 2001, 19, 765–768. [Google Scholar] [CrossRef]
  41. Zhang, Y.-M.; Zhang, H.-M.; Liu, Z.-H.; Li, H.-C.; Guo, X.-L.; Li, G.-L. The wheat NHX antiporter gene TaNHX2 confers salt tolerance in transgenic alfalfa by increasing the retention capacity of intracellular potassium. Plant Mol. Biol. 2015, 87, 317–327. [Google Scholar] [CrossRef]
  42. Zhang, H.-B.; Li, Y.; Wang, B.; Chee, P.W. Recent advances in cotton genomics. Int. J. Plant Genom. 2008, 2008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Yang, Z.; Zhang, C.; Yang, X.; Liu, K.; Wu, Z.; Zhang, X.; Zheng, W.; Xun, Q.; Liu, C.; Lu, L. PAG1, a cotton brassinosteroid catabolism gene, modulates fiber elongation. New Phytol. 2014, 203, 437–448. [Google Scholar] [CrossRef] [PubMed]
  44. Huang, J.; Pang, C.; Fan, S.; Song, M.; Yu, J.; Wei, H.; Ma, Q.; Li, L.; Zhang, C.; Yu, S. Genome-wide analysis of the family 1 glycosyltransferases in cotton. Mol. Genet. Genom. 2015, 290, 1805–1818. [Google Scholar] [CrossRef]
  45. Ahmad, S.; Khan, N.; Iqbal, M.Z.; Hussain, A.; Hassan, M. Salt tolerance of cotton (Gossypium hirsutum L.). Asian J. Plant Sci. 2002, 1, 715–719. [Google Scholar]
  46. Zhang, G.-W.; Lu, H.-L.; Zhang, L.; Chen, B.-L.; Zhou, Z.-G. Salt tolerance evaluation of cotton (Gossypium hirsutum) at its germinating and seedling stages and selection of related indices. Yingyong Shengtai Xuebao 2011, 22, 2045–2053. [Google Scholar] [PubMed]
  47. Shaheen, H.L.; Iqbal, M.; Azeem, M.; Shahbaz, M.; Shehzadi, M. K-priming positively modulates growth and nutrient status of salt-stressed cotton (Gossypium hirsutum) seedlings. Arch. Agron. Soil Sci. 2016, 62, 759–768. [Google Scholar] [CrossRef]
  48. Silberbush, M.; Ben-Asher, J. The effect of salinity on parameters of potassium and nitrate uptake of cotton. Commun. Soil Sci. Plan. 1987, 18, 65–81. [Google Scholar] [CrossRef]
  49. Bernstein, L.; Hayward, H. Physiology of salt tolerance. Annu. Rev. Plant Physiol. 1958, 9, 25–46. [Google Scholar] [CrossRef]
  50. Peng, J.; Liu, J.; Zhang, L.; Luo, J.; Dong, H.; Ma, Y.; Zhao, X.; Chen, B.; Sui, N.; Zhou, Z. Effects of soil salinity on sucrose metabolism in cotton leaves. PLoS ONE 2016, 11, e0156241. [Google Scholar] [CrossRef]
  51. Longenecker, D. The influence of high sodium in soils upon fruiting and shedding, boll characteristics, fiber properties, and yields of two cotton species. Soil Sci. 1974, 118, 387–396. [Google Scholar] [CrossRef]
  52. Satir, O.; Berberoglu, S. Crop yield prediction under soil salinity using satellite derived vegetation indices. Field Crops Res. 2016, 192, 134–143. [Google Scholar] [CrossRef]
  53. Li, F.; Fan, G.; Lu, C.; Xiao, G.; Zou, C.; Kohel, R.J.; Ma, Z.; Shang, H.; Ma, X.; Wu, J. Genome sequence of cultivated Upland cotton (Gossypium hirsutum TM-1) provides insights into genome evolution. Nat. Biotechnol. 2015, 33, 524–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Zhang, T.; Hu, Y.; Jiang, W.; Fang, L.; Guan, X.; Chen, J.; Zhang, J.; Saski, C.A.; Scheffler, B.E.; Stelly, D.M. Sequencing of allotetraploid cotton (Gossypium hirsutum L. acc. TM-1) provides a resource for fiber improvement. Nat. Biotechnol. 2015, 33, 531–537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Li, F.; Fan, G.; Wang, K.; Sun, F.; Yuan, Y.; Song, G.; Li, Q.; Ma, Z.; Lu, C.; Zou, C. Genome sequence of the cultivated cotton Gossypium arboreum. Nat. Genet. 2014, 46, 567–572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Du, X.; Huang, G.; He, S.; Yang, Z.; Sun, G.; Ma, X.; Li, N.; Zhang, X.; Sun, J.; Liu, M. Resequencing of 243 diploid cotton accessions based on an updated A genome identifies the genetic basis of key agronomic traits. Nat. Genet. 2018, 50, 796–802. [Google Scholar] [CrossRef]
  57. Wang, K.; Wang, Z.; Li, F.; Ye, W.; Wang, J.; Song, G.; Yue, Z.; Cong, L.; Shang, H.; Zhu, S. The draft genome of a diploid cotton Gossypium raimondii. Nat. Genet. 2012, 44, 1098–1103. [Google Scholar] [CrossRef]
  58. Liu, X.; Zhao, B.; Zheng, H.-J.; Hu, Y.; Lu, G.; Yang, C.-Q.; Chen, J.-D.; Chen, J.-J.; Chen, D.-Y.; Zhang, L. Gossypium barbadense genome sequence provides insight into the evolution of extra-long staple fiber and specialized metabolites. Sci. Rep. 2015, 5, 14139. [Google Scholar] [CrossRef]
  59. Finn, R.D.; Coggill, P.; Eberhardt, R.Y.; Eddy, S.R.; Mistry, J.; Mitchell, A.L.; Potter, S.C.; Punta, M.; Qureshi, M.; Sangrador-Vegas, A. The Pfam protein families database: Towards a more sustainable future. Nucleic Acids Res. 2015, 44, D279–D285. [Google Scholar] [CrossRef]
  60. Zhu, T.; Liang, C.; Meng, Z.; Sun, G.; Meng, Z.; Guo, S.; Zhang, R. CottonFGD: An integrated functional genomics database for cotton. BMC Plant Biol. 2017, 17, 1–9. [Google Scholar] [CrossRef] [Green Version]
  61. Finn, R.D.; Clements, J.; Eddy, S.R. HMMER web server: Interactive sequence similarity searching. Nucleic Acids Res. 2011, 39, W29–W37. [Google Scholar] [CrossRef] [Green Version]
  62. Moller, S.; Croning, M.D.; Apweiler, R. Evaluation of methods for the prediction of membrane spanning regions. Bioinformatics 2001, 17, 646–653. [Google Scholar] [CrossRef] [Green Version]
  63. Yu, C.S.; Lin, C.J.; Hwang, J.K. Predicting subcellular localization of proteins for Gram-negative bacteria by support vector machines based on n-peptide compositions. Protein Sci. 2004, 13, 1402–1406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Yu, C.S.; Chen, Y.C.; Lu, C.H.; Hwang, J.K. Prediction of protein subcellular localization. Proteins: Struct. Funct. Bioinform. 2006, 64, 643–651. [Google Scholar] [CrossRef]
  65. Blom, N.; Gammeltoft, S.; Brunak, S. Sequence and structure-based prediction of eukaryotic protein phosphorylation sites. J. Mol. Biol. 1999, 294, 1351–1362. [Google Scholar] [CrossRef] [PubMed]
  66. Bailey, T.L.; Boden, M.; Buske, F.A.; Frith, M.; Grant, C.E.; Clementi, L.; Ren, J.; Li, W.W.; Noble, W.S. MEME SUITE: Tools for motif discovery and searching. Nucleic Acids Res. 2009, 37, W202–W208. [Google Scholar] [CrossRef] [PubMed]
  67. Hima Kumari, P.; Anil Kumar, S.; Ramesh, K.; Sudhakar Reddy, P.; Nagaraju, M.; Bhanu Prakash, A.; Shah, T.; Henderson, A.; Srivastava, R.K.; Rajasheker, G. Genome-wide identification and analysis of Arabidopsis sodium proton antiporter (NHX) and human sodium proton exchanger (NHE) homologs in sorghum bicolor. Genes 2018, 9, 236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Taji, T.; Sakurai, T.; Mochida, K.; Ishiwata, A.; Kurotani, A.; Totoki, Y.; Toyoda, A.; Sakaki, Y.; Seki, M.; Ono, H. Large-scale collection and annotation of full-length enriched cDNAs from a model halophyte, Thellungiella halophila. BMC Plant Biol. 2008, 8, 115. [Google Scholar] [CrossRef] [Green Version]
  69. Tian, F.; Chang, E.; Li, Y.; Sun, P.; Hu, J.; Zhang, J. Expression and integrated network analyses revealed functional divergence of NHX-type Na+/H+ exchanger genes in poplar. Sci. Rep. 2017, 7, 1–17. [Google Scholar] [CrossRef] [Green Version]
  70. Sandhu, D.; Pudussery, M.V.; Kaundal, R.; Suarez, D.L.; Kaundal, A.; Sekhon, R.S. Molecular characterization and expression analysis of the Na+/H+ exchanger gene family in Medicago truncatula. Funct. Integr. Genom. 2018, 18, 141–153. [Google Scholar] [CrossRef]
  71. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  72. Letunic, I.; Bork, P. Interactive Tree Of Life (iTOL) v4: Recent updates and new developments. Nucleic Acids Res. 2019, 47, W256–W259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools—An integrative toolkit developed for interactive analyses of big biological data. Mol. Plant 2020. [Google Scholar] [CrossRef] [PubMed]
  74. Suyama, M.; Torrents, D.; Bork, P. PAL2NAL: Robust conversion of protein sequence alignments into the corresponding codon alignments. Nucleic Acids Res. 2006, 34, W609–W612. [Google Scholar] [CrossRef] [Green Version]
  75. Lescot, M.; Déhais, P.; Thijs, G.; Marchal, K.; Moreau, Y.; Van de Peer, Y.; Rouzé, P.; Rombauts, S. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002, 30, 325–327. [Google Scholar] [CrossRef] [PubMed]
  76. Guo, A.; Zhu, Q.; Chen, X.; Luo, J. GSDS: A gene structure display server. Yi Chuan = Hereditas 2007, 29, 1023–1026. [Google Scholar] [CrossRef]
  77. Artico, S.; Nardeli, S.M.; Brilhante, O.; Grossi-de-Sa, M.F.; Alves-Ferreira, M. Identification and evaluation of new reference genes in Gossypium hirsutum for accurate normalization of real-time quantitative RT-PCR data. BMC Plant Biol. 2010, 10, 49. [Google Scholar] [CrossRef] [Green Version]
  78. Paterson, A.H.; Wendel, J.F.; Gundlach, H.; Guo, H.; Jenkins, J.; Jin, D.; Llewellyn, D.; Showmaker, K.C.; Shu, S.; Udall, J. Repeated polyploidization of Gossypium genomes and the evolution of spinnable cotton fibres. Nature 2012, 492, 423–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Eulgem, T.; Rushton, P.J.; Schmelzer, E.; Hahlbrock, K.; Somssich, I.E. Early nuclear events in plant defence signalling: Rapid gene activation by WRKY transcription factors. EMBO J. 1999, 18, 4689–4699. [Google Scholar] [CrossRef] [Green Version]
  80. Abul-Naas, A.A.; Omran, M.S. Salt tolerance of seventeen cotton cultivars during germination and early seedling development. Z Acker Pflanzenbau 1975, 140, 229–236. [Google Scholar]
  81. Ashraf, M. Salt tolerance of cotton: Some new advances. Crit. Rev. Plant Sci. 2002, 21, 1–30. [Google Scholar] [CrossRef]
  82. Witt, T.W.; Ulloa, M.; Schwartz, R.C.; Ritchie, G.L. Response to deficit irrigation of morphological, yield and fiber quality traits of upland (Gossypium hirsutum L.) and Pima (G. barbadense L.) cotton in the Texas High Plains. Field Crop. Res. 2020, 249, 107759. [Google Scholar] [CrossRef]
  83. Hu, Y.; Chen, J.; Fang, L.; Zhang, Z.; Ma, W.; Niu, Y.; Ju, L.; Deng, J.; Zhao, T.; Lian, J. Gossypium barbadense and Gossypium hirsutum genomes provide insights into the origin and evolution of allotetraploid cotton. Nat. Genet. 2019, 51, 739–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Orlowski, J.; Grinstein, S. Diversity of the mammalian sodium/proton exchanger SLC9 gene family. Pflug. Arch. 2004, 447, 549–565. [Google Scholar] [CrossRef] [PubMed]
  85. Shi, H.; Quintero, F.J.; Pardo, J.M.; Zhu, J.K. The putative plasma membrane Na(+)/H(+) antiporter SOS1 controls long-distance Na(+) transport in plants. Plant Cell 2002, 14, 465–477. [Google Scholar] [CrossRef] [Green Version]
  86. Wu, G.-Q.; Wang, J.-L.; Li, S.-J. Genome-wide identification of Na+/H+ antiporter (NHX) genes in sugar beet (Beta vulgaris L.) and their regulated expression under salt stress. Genes 2019, 10, 401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Kinsella, J.L.; Aronson, P.S. Amiloride inhibition of the Na+-H+ exchanger in renal microvillus membrane vesicles. Am. J. Physiol. 1981, 241, F374–F379. [Google Scholar] [CrossRef]
  88. Blumwald, E.; Poole, R.J. Salt tolerance in suspension cultures of sugar beet: Induction of na/h antiport activity at the tonoplast by growth in salt. Plant Physiol. 1987, 83, 884–887. [Google Scholar] [CrossRef] [Green Version]
  89. Counillon, L.; Franchi, A.; Pouyssegur, J. A point mutation of the Na+/H+ exchanger gene (NHE1) and amplification of the mutated allele confer amiloride resistance upon chronic acidosis. Proc. Natl. Acad. Sci. USA 1993, 90, 4508–4512. [Google Scholar] [CrossRef] [Green Version]
  90. Rong, J.; Feltus, F.A.; Liu, L.; Lin, L.; Paterson, A.H. Gene copy number evolution during tetraploid cotton radiation. Heredity 2010, 105, 463–472. [Google Scholar] [CrossRef] [Green Version]
  91. Conant, G.C.; Wolfe, K.H. Turning a hobby into a job: How duplicated genes find new functions. Nat Rev. Genet. 2008, 9, 938–950. [Google Scholar] [CrossRef]
  92. Wang, Y.; Hou, J.; Liu, H.; Li, T.; Wang, K.; Hao, C.; Liu, H.; Zhang, X. TaBT1, affecting starch synthesis and thousand kernel weight, underwent strong selection during wheat improvement. J. Exp. Bot. 2019, 70, 1497–1511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Bowers, J.E.; Chapman, B.A.; Rong, J.; Paterson, A.H. Unravelling angiosperm genome evolution by phylogenetic analysis of chromosomal duplication events. Nature 2003, 422, 433–438. [Google Scholar] [CrossRef] [PubMed]
  94. Martin, C.; Paz-Ares, J. MYB transcription factors in plants. Trends Genet. 1997, 13, 67–73. [Google Scholar] [CrossRef]
  95. Liu, L.; Xu, W.; Hu, X.; Liu, H.; Lin, Y. W-box and G-box elements play important roles in early senescence of rice flag leaf. Sci. Rep. 2016, 6, 20881. [Google Scholar] [CrossRef] [Green Version]
  96. Venema, K.; Quintero, F.J.; Pardo, J.M.; Donaire, J.P. The Arabidopsis Na+/H+ exchanger AtNHX1 catalyzes low affinity Na+ and K+ transport in reconstituted liposomes. J. Biol. Chem. 2002, 277, 2413–2418. [Google Scholar] [CrossRef] [Green Version]
  97. Xia, T.; Apse, M.P.; Aharon, G.S.; Blumwald, E. Identification and characterization of a NaCl-inducible vacuolar Na+/H+ antiporter in Beta vulgaris. Physiol. Plant. 2002, 116, 206–212. [Google Scholar] [CrossRef]
  98. Rodríguez-Rosales, M.P.; Jiang, X.; Gálvez, F.J.; Aranda, M.N.; Cubero, B.; Venema, K. Overexpression of the tomato K+/H+ antiporter LeNHX2 confers salt tolerance by improving potassium compartmentalization. New Phytol. 2008, 179, 366–377. [Google Scholar] [CrossRef]
  99. Vasekina, A.; Yershov, P.; Reshetova, O.; Tikhonova, T.; Lunin, V.; Trofimova, M.; Babakov, A. Vacuolar Na+/H+ antiporter from barley: Identification and response to salt stress. Biochemistry (Moscow) 2005, 70, 100–107. [Google Scholar] [CrossRef]
  100. Zörb, C.; Noll, A.; Karl, S.; Leib, K.; Yan, F.; Schubert, S. Molecular characterization of Na+/H+ antiporters (ZmNHX) of maize (Zea mays L.) and their expression under salt stress. J. Plant Physiol. 2005, 162, 55–66. [Google Scholar] [CrossRef]
  101. Brini, F.; Gaxiola, R.A.; Berkowitz, G.A.; Masmoudi, K. Cloning and characterization of a wheat vacuolar cation/proton antiporter and pyrophosphatase proton pump. Plant Physiol. Biochem. 2005, 43, 347–354. [Google Scholar] [CrossRef]
  102. Li, W.Y.F.; Wong, F.L.; Tsai, S.N.; Phang, T.H.; Shao, G.; Lam, H.M. Tonoplast-located GmCLC1 and GmNHX1 from soybean enhance NaCl tolerance in transgenic bright yellow (BY)-2 cells. Plant Cell Environ. 2006, 29, 1122–1137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Zeng, Y.; Li, Q.; Wang, H.; Zhang, J.; Du, J.; Feng, H.; Blumwald, E.; Yu, L.; Xu, G. Two NHX-type transporters from Helianthus tuberosus improve the tolerance of rice to salinity and nutrient deficiency stress. Plant Biotechnol. J. 2018, 16, 310–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Fukuda, A.; Nakamura, A.; Hara, N.; Toki, S.; Tanaka, Y. Molecular and functional analyses of rice NHX-type Na+/H+ antiporter genes. Planta 2011, 233, 175–188. [Google Scholar] [CrossRef] [PubMed]
  105. Ma, Y.; Wang, J.; Zhong, Y.; Geng, F.; Cramer, G.R.; Cheng, Z.-M.M. Subfunctionalization of cation/proton antiporter 1 genes in grapevine in response to salt stress in different organs. Hortic. Res. 2015, 2, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Bao, A.K.; Du, B.Q.; Touil, L.; Kang, P.; Wang, Q.L.; Wang, S.M. Co-expression of tonoplast Cation/H+ antiporter and H+-pyrophosphatase from xerophyte Zygophyllum xanthoxylum improves alfalfa plant growth under salinity, drought and field conditions. Plant Biotechnol. J. 2016, 14, 964–975. [Google Scholar] [CrossRef]
  107. Wang, B.; Zhai, H.; He, S.; Zhang, H.; Ren, Z.; Zhang, D.; Liu, Q. A vacuolar Na+/H+ antiporter gene, IbNHX2, enhances salt and drought tolerance in transgenic sweetpotato. Sci. Hortic. 2016, 201, 153–166. [Google Scholar] [CrossRef]
  108. Lu, W.; Guo, C.; Li, X.; Duan, W.; Ma, C.; Zhao, M.; Gu, J.; Du, X.; Liu, Z.; Xiao, K. Overexpression of TaNHX3, a vacuolar Na+/H+ antiporter gene in wheat, enhances salt stress tolerance in tobacco by improving related physiological processes. Plant Physiol. Biochem. 2014, 76, 17–28. [Google Scholar] [CrossRef]
  109. Blumwald, E. Sodium transport and salt tolerance in plants. Curr. Opin. Cell Biol. 2000, 12, 431–434. [Google Scholar] [CrossRef]
  110. Yadav, N.S.; Shukla, P.S.; Jha, A.; Agarwal, P.K.; Jha, B. The SbSOS1 gene from the extreme halophyte Salicornia brachiata enhances Na+ loading in xylem and confers salt tolerance in transgenic tobacco. BMC Plant Biol. 2012, 12, 188. [Google Scholar] [CrossRef] [Green Version]
  111. Zhang, W.-D.; Wang, P.; Bao, Z.; Ma, Q.; Duan, L.-J.; Bao, A.-K.; Zhang, J.-L.; Wang, S.-M. SOS1, HKT1; 5, and NHX1 synergistically modulate Na+ homeostasis in the halophytic grass Puccinellia tenuiflora. Front. Plant Sci. 2017, 8, 576. [Google Scholar]
  112. Ma, Q.; Li, Y.-X.; Yuan, H.-J.; Hu, J.; Wei, L.; Bao, A.-K.; Zhang, J.-L.; Wang, S.-M. ZxSOS1 is essential for long-distance transport and spatial distribution of Na+ and K+ in the xerophyte Zygophyllum xanthoxylum. Plant Soil 2014, 374, 661–676. [Google Scholar] [CrossRef]
  113. Quintero, F.J.; Martinez-Atienza, J.; Villalta, I.; Jiang, X.; Kim, W.-Y.; Ali, Z.; Fujii, H.; Mendoza, I.; Yun, D.-J.; Zhu, J.-K. Activation of the plasma membrane Na/H antiporter Salt-Overly-Sensitive 1 (SOS1) by phosphorylation of an auto-inhibitory C-terminal domain. Proc. Natl. Acad. Sci. USA 2011, 108, 2611–2616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Katiyar-Agarwal, S.; Zhu, J.; Kim, K.; Agarwal, M.; Fu, X.; Huang, A.; Zhu, J.-K. The plasma membrane Na+/H+ antiporter SOS1 interacts with RCD1 and functions in oxidative stress tolerance in Arabidopsis. Proc. Natl. Acad. Sci. USA 2006, 103, 18816–18821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Weinl, S.; Kudla, J. The CBL–CIPK Ca2+-decoding signaling network: Function and perspectives. New Phytol. 2009, 184, 517–528. [Google Scholar] [CrossRef] [PubMed]
  116. Kim, W.-Y.; Ali, Z.; Park, H.J.; Park, S.J.; Cha, J.-Y.; Perez-Hormaeche, J.; Quintero, F.J.; Shin, G.; Kim, M.R.; Qiang, Z. Release of SOS2 kinase from sequestration with GIGANTEA determines salt tolerance in Arabidopsis. Nat. Commun. 2013, 4, 1–13. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Gene Structure of the G. barbadense NHX transporters. Red box represents the exons and the black lines represent the introns.
Figure 1. Gene Structure of the G. barbadense NHX transporters. Red box represents the exons and the black lines represent the introns.
Genes 11 00803 g001
Figure 2. Phylogenetic tree of sodium transporters between the NHX transporters of 11 plant species by the neighbor-end joining method using MEGA 10.0. The tree divides all the 125 NHX genes into three groups based on their subcellular localization. Prefixes such as Gh, Gb, Gr, Ga, At, Vv, Ptr, Sb, Medtr, Eh, and Pp were used before the name of the species G. hirsutum, G. barbadense, G. raimondii, G. arboreum, A. thaliana, V. vinifera, P. trichocarpa, S. bicolor, M. truncatula, E. halophilum, and P. patens, respectively. G. barbadense genes are represented by bold letters. The amino acid sequences used in phylogenetic analysis are provided in Table S1.
Figure 2. Phylogenetic tree of sodium transporters between the NHX transporters of 11 plant species by the neighbor-end joining method using MEGA 10.0. The tree divides all the 125 NHX genes into three groups based on their subcellular localization. Prefixes such as Gh, Gb, Gr, Ga, At, Vv, Ptr, Sb, Medtr, Eh, and Pp were used before the name of the species G. hirsutum, G. barbadense, G. raimondii, G. arboreum, A. thaliana, V. vinifera, P. trichocarpa, S. bicolor, M. truncatula, E. halophilum, and P. patens, respectively. G. barbadense genes are represented by bold letters. The amino acid sequences used in phylogenetic analysis are provided in Table S1.
Genes 11 00803 g002
Figure 3. Conserved motif analysis of the NHX genes. (a) Motifs of the GbNHX Amiloride binding site) are represented by motif 7. (b) Motifs of the GhNHX Amiloride binding site are shown by motif 1. The sequence for amiloride binding site ((L/F)FF(I/L)(Y/F)LLPPI is highlighted.
Figure 3. Conserved motif analysis of the NHX genes. (a) Motifs of the GbNHX Amiloride binding site) are represented by motif 7. (b) Motifs of the GhNHX Amiloride binding site are shown by motif 1. The sequence for amiloride binding site ((L/F)FF(I/L)(Y/F)LLPPI is highlighted.
Genes 11 00803 g003
Figure 4. Collinearity analysis of G. barbadense (a) and G. hirsutum (b) (A and D) orthologs in the G. arboreum (A Chr) and G. raimondii (D Chr) genomes. Orthologs are connected by colored lines.
Figure 4. Collinearity analysis of G. barbadense (a) and G. hirsutum (b) (A and D) orthologs in the G. arboreum (A Chr) and G. raimondii (D Chr) genomes. Orthologs are connected by colored lines.
Genes 11 00803 g004
Figure 5. Cis-elements of the NHX transporters. Vertical axis represents the number of Cis-elements, and the horizontal axis shows the genes name.
Figure 5. Cis-elements of the NHX transporters. Vertical axis represents the number of Cis-elements, and the horizontal axis shows the genes name.
Genes 11 00803 g005
Figure 6. Relative expression level analysis of NHX in Gossypium Hirsutum. Relative expression of different NHXs is shown under the controlled conditions and salinity stress in different tissues at different time intervals. Y-axis shows the gene names and X-axis represents the tissue and time interval. Colors represent the expression level normalized against the control tissues. LC: leaf control; RC: root control; SC: stem Control; LT: treated leaf; RT: treated root; ST: treated stem.
Figure 6. Relative expression level analysis of NHX in Gossypium Hirsutum. Relative expression of different NHXs is shown under the controlled conditions and salinity stress in different tissues at different time intervals. Y-axis shows the gene names and X-axis represents the tissue and time interval. Colors represent the expression level normalized against the control tissues. LC: leaf control; RC: root control; SC: stem Control; LT: treated leaf; RT: treated root; ST: treated stem.
Genes 11 00803 g006
Figure 7. Relative expressions of 10 GbNHX genes under salinity stress based on a qRT-PCR. The values are the means ± standard deviations (SD) of three replicates. Gene specific primers list is provided in Table S3.
Figure 7. Relative expressions of 10 GbNHX genes under salinity stress based on a qRT-PCR. The values are the means ± standard deviations (SD) of three replicates. Gene specific primers list is provided in Table S3.
Genes 11 00803 g007
Figure 8. String analysis of GbNHX interacting proteins.
Figure 8. String analysis of GbNHX interacting proteins.
Genes 11 00803 g008
Figure 9. Gene ontology (GO) terms of GbNHX genes. Different GO terms are represented by different shapes.
Figure 9. Gene ontology (GO) terms of GbNHX genes. Different GO terms are represented by different shapes.
Genes 11 00803 g009
Table 1. Characteristics of the G. barbadense sodium-proton antiporter (NHX) transporters.
Table 1. Characteristics of the G. barbadense sodium-proton antiporter (NHX) transporters.
Gene NameGene IDProtein (aa)CDS (bp)MW (kDa)pILocalizationNa+/H+ Exchanger Domain (Start–End)
Gb-NHX1Gbar_D07G012760164145518.66610.438VacNil
Gb-NHX2-1AGbar_A02G004630535152759.0638.663Vac22–437
Gb-NHX2-2AGbar_A08G023430320160835.54810.181Vac7–212
Gb-NHX2-3AGbar_A09G007060542345959.8447.858Vac29–444
Gb-NHX2-4AGbar_A09G025000541158759.7219.05Vac25–445
Gb-NHX2-5AGbar_A11G02801054096359.5866.989Vac16–440
Gb-NHX2-6AGbar_A11G025170543162960.1067.657Vac24–445
Gb-NHX2-7AGbar_A12G000720445162649.3718.063Vac17–420
Gb-NHX2-8AGbar_A13G011300541163259.7068.552Vac25–444
Gb-NHX4-1AGbar_A01G007690508162356.9777.591Vac16–426
Gb-NHX6-1AGbar_A01G002880484133853.2466.795Endo25–433
Gb-NHX6-2AGbar_A06G019530528162658.415.514Vac25–437
Gb-NHX7-1AGbar_A03G01287011521584128.076.878PM29–445
Gb-NHX2-1DGbar_D08G024100551157861.3158.457Vac30–434
GB-NHX2-2DGbar_D02G005160535160859.1788.453Vac22–437
Gb-NHX2-3DGbar_D09G006790542345960.0158.731Vac29–444
Gb-NHX2-4DGbar_D09G024630541157259.7059.175Vac29–444
Gb-NHX2-5DGbar_D11G02610049749555.1197.009Vac4–406
Gb-NHX2-6DGbar_D11G028500542165659.7586.42Vac19–448
Gb-NHX2-7DGbar_D12G000860525162958.1268.549Vac25–444
Gb-NHX2-8DGbar_D13G011070541162659.7158.554Vac31–442
Gb-NHX4-1DGbar_D01G007950525149459.1397.62Vac21–441
GB-NHX6-1DGbar_D01G003050527162958.0565.978Endo28–437
Gb-NHX6-2DGbar_D06G020390523157857.725.494Endo28–432
Gb-NHX7-1DGbar_D02G01481011521626128.146.764PM31–443
aa: amino acid; pI: isoelectric point; MW: molecular weight; Vac: vacuole; Pm: plasma membrane; Endo: endomembrane.
Table 2. Chromosomal location of the NHX genes in the Gossypium species.
Table 2. Chromosomal location of the NHX genes in the Gossypium species.
ChromosomeG. arboreumG. raimondiiG. barbadenseG. hirsutum
A01Ga_NHX6-1 Gb-NHX6-1AGh_NHX6-1A
A01Ga_NHX4 Gb-NHX4-1AGh_NHX4-1A
A02 Gb-NHX2-1AGh_NHX2-1A
A03Ga_NHX2-1
A03Ga_NHX7 Gb-NHX7-1AGh_NHX7-1A
A06Ga_NHX1
A06Ga_NHX6-2 Gb-NHX6-2AGh_NHX6-2A
A08Ga_NHX2-2 Gb-NHX2-2A
A09Ga_NHX2-3 Gb-NHX2-3AGh_NHX2-2A
A09Ga_NHX2-4 Gb-NHX2-4AGh_NHX2-3A
A11Ga_NHX2-5 Gb-NHX2-6AGh_NHX2-4A
A11Ga_NHX2-6 Gb-NHX2-5AGh_NHX2-5A
A12Ga_NHX2-7 Gb-NHX2-7AGh_NHX2-6A
A13Ga_NHX2-8 Gb-NHX2-8AGh_NHX2-7A
D01 Gb-NHX6-1DGh_NHX6-1D
D01 Gb-NHX4-1DGh_NHX4-1D
D02 Gr_NHX6-1Gb-NHX2-2DGh_NHX2-1D
D02 Gr_NHX4Gb-NHX7-1DGh_NHX7-1D
D04 Gr_NHX2-1
D05 Gr_NHX2-2
D05 Gr_NHX7
D06 Gr_NHX2-3Gb-NHX6-2DGh_NHX6-2D
D06 Gr_NHX2-4
D07 Gr_NHX2-5Gb-NHX1Gh_NHX1
D07 Gr_NHX2-6
D08 Gr_NHX2-7Gb-NHX2-1D
D09 Gb-NHX2-3DGh_NHX2-2D
D09 Gb-NHX2-4DGh_NHX2-3D
D10 Gr_NHX2-8
D10 Gr_NHX6-2
D11 Gb-NHX2-5DGh_NHX2-4D
D11 Gb-NHX2-6DGh_NHX2-5D
D12 Gb-NHX2-7DGh_NHX2-6D
D13 Gr_NHX2-9Gb-NHX2-8DGh_NHX2-7D
Table 3. Orthologous and paralogous gene pairs for Gb and Gh.
Table 3. Orthologous and paralogous gene pairs for Gb and Gh.
G. barbadense Orthologous/ParalogousG. hirsutum Orthologous/Paralogous
Gene IDGene IDGene IDGene ID
Gb-NHX6-1AGB-NHX6-1DGh_NHX6-1Gh_NHX2-2
Gb-NHX4-1AGb-NHX4-1DGh_NHX6-1Gh_NHX2-3
Gb-NHX2-1AGb-NHX2-4AGh_NHX6-1Gh_NHX6-3
Gb-NHX2-1AGb-NHX2-3AGh_NHX6-1Gh_NHX2-9
Gb-NHX2-1AGB-NHX2-2DGh_NHX6-1Gh_NHX2-10
Gb-NHX2-1AGb-NHX2-4DGh_NHX4-1Gh_NHX4-2
Gb-NHX2-1AGb-NHX2-3DGh_NHX2-1Gh_NHX2-2
Gb-NHX7-1AGb-NHX7-1DGh_NHX2-1Gh_NHX2-3
Gb-NHX6-2AGb-NHX6-2DGh_NHX2-1Gh_NHX2-6
Gb-NHX2-2AGb-NHX2-6AGh_NHX2-1Gh_NHX2-8
Gb-NHX2-2AGb-NHX2-1D.1Gh_NHX2-1Gh_NHX2-10
Gb-NHX2-2AGb-NHX2-8D.1Gh_NHX2-1Gh_NHX2-9
Gb-NHX2-4AGB-NHX2-2DGh_NHX7-1Gh_NHX7-2
Gb-NHX2-3AGB-NHX2-2DGh_NHX2-2Gh_NHX2-3
Gb-NHX2-4AGb-NHX2-4DGh_NHX2-2Gh_NHX2-8
Gb-NHX2-3AGb-NHX2-3DGh_NHX2-2Gh_NHX6-4
Gb-NHX2-6AGb-NHX2-8AGh_NHX2-2Gh_NHX2-9
Gb-NHX2-6AGb-NHX2-1DGh_NHX2-3Gh_NHX2-8
Gb-NHX2-6AGb-NHX2-5DGh_NHX2-3Gh_NHX2-10
Gb-NHX2-5AGb-NHX2-6DGh_NHX2-3Gh_NHX2-9
Gb-NHX2-6AGb-NHX2-8DGh_NHX2-4Gh_NHX2-7
Gb-NHX2-7AGb-NHX2-7DGh_NHX2-4Gh_NHX2-11
Gb-NHX2-8AGb-NHX2-1DGh_NHX2-4Gh_NHX2-14
Gb-NHX2-8AGb-NHX2-5DGh_NHX2-5Gh_NHX2-12
Gb-NHX2-8AGb-NHX2-8DGh_NHX2-6Gh_NHX2-8
GB-NHX2-2DGb-NHX2-4DGh_NHX2-7Gh_NHX2-11
GB-NHX2-2DGb-NHX2-3DGh_NHX2-7Gh_NHX2-14
Gb-NHX2-1DGb-NHX2-5DGh_NHX2-8Gh_NHX2-10
Gb-NHX2-1DGb-NHX2-8DGh_NHX2-8Gh_NHX2-9
Gb-NHX2-5DGb-NHX2-8DGh_NHX6-4Gh_NHX2-9
Gh_NHX2-11Gh_NHX2-14

Share and Cite

MDPI and ACS Style

Akram, U.; Song, Y.; Liang, C.; Abid, M.A.; Askari, M.; Myat, A.A.; Abbas, M.; Malik, W.; Ali, Z.; Guo, S.; et al. Genome-Wide Characterization and Expression Analysis of NHX Gene Family under Salinity Stress in Gossypium barbadense and Its Comparison with Gossypium hirsutum. Genes 2020, 11, 803. https://doi.org/10.3390/genes11070803

AMA Style

Akram U, Song Y, Liang C, Abid MA, Askari M, Myat AA, Abbas M, Malik W, Ali Z, Guo S, et al. Genome-Wide Characterization and Expression Analysis of NHX Gene Family under Salinity Stress in Gossypium barbadense and Its Comparison with Gossypium hirsutum. Genes. 2020; 11(7):803. https://doi.org/10.3390/genes11070803

Chicago/Turabian Style

Akram, Umar, Yuhan Song, Chengzhen Liang, Muhammad Ali Abid, Muhammad Askari, Aye Aye Myat, Mubashir Abbas, Waqas Malik, Zulfiqar Ali, Sandui Guo, and et al. 2020. "Genome-Wide Characterization and Expression Analysis of NHX Gene Family under Salinity Stress in Gossypium barbadense and Its Comparison with Gossypium hirsutum" Genes 11, no. 7: 803. https://doi.org/10.3390/genes11070803

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop